Вы находитесь на странице: 1из 14

6896 The Journal of Neuroscience, May 14, 2014 34(20):6896 6909

Behavioral/Cognitive

Optogenetic Manipulation of Activity and Temporally


Controlled Cell-Specific Ablation Reveal a Role for MCH
Neurons in Sleep/Wake Regulation
Tomomi Tsunematsu,1,2 Takafumi Ueno,3 Sawako Tabuchi,1,2,4 Ayumu Inutsuka,1 Kenji F. Tanaka,5 Hidetoshi Hasuwa,6
Thomas S. Kilduff,7 Akira Terao3, and Akihiro Yamanaka1
1Department of Neuroscience II, Research Institute of Environmental Medicine, Nagoya University, Nagoya 464-8601, Japan, 2The Japan Society for the
Promotion of Sciences, Tokyo 102-8472, Japan, 3Laboratory of Biochemistry, Department of Biomedical Sciences, Graduate School of Veterinary Medicine,
Hokkaido University, Sapporo 060-0818, Japan, 4Department of Physiological Sciences, The Graduate University for Advanced Studies, Okazaki 444-8787,
Japan, 5Department of Neuropsychiatry, School of Medicine, Keio University, Tokyo 160-8582, Japan, 6Research Institute for Microbial Disease, Osaka
University, Suita 565-0781 Japan, and 7Center for Neuroscience, Biosciences Division, SRI International, Menlo Park, California 94025

Melanin-concentrating hormone (MCH) is a neuropeptide produced in neurons sparsely distributed in the lateral hypothalamic area.
Recent studies have reported that MCH neurons are active during rapid eye movement (REM) sleep, but their physiological role in the
regulation of sleep/wakefulness is not fully understood. To determine the physiological role of MCH neurons, newly developed transgenic
mouse strains that enable manipulation of the activity and fate of MCH neurons in vivo were generated using the recently developed
knockin-mediated enhanced gene expression by improved tetracycline-controlled gene induction system. The activity of these cells was
controlled by optogenetics by expressing channelrhodopsin2 (E123T/T159C) or archaerhodopsin-T in MCH neurons. Acute optogenetic
activation of MCH neurons at 10 Hz induced transitions from non-REM (NREM) to REM sleep and increased REM sleep time in conjunction with decreased NREM sleep. Activation of MCH neurons while mice were in NREM sleep induced REM sleep, but activation during
wakefulness was ineffective. Acute optogenetic silencing of MCH neurons using archaerhodopsin-T had no effect on any vigilance states.
Temporally controlled ablation of MCH neurons by cell-specific expression of diphtheria toxin A increased wakefulness and decreased
NREM sleep duration without affecting REM sleep. Together, these results indicate that acute activation of MCH neurons is sufficient, but
not necessary, to trigger the transition from NREM to REM sleep and that MCH neurons also play a role in the initiation and maintenance
of NREM sleep.
Key words: ablation; cell fate; channelrhodopsin2; hypothalamus; optogenetics; REM sleep

Introduction
Melanin-concentrating hormone (MCH)-producing neurons
are coexpressed with orexin/hypocretin-producing neurons in
Received Dec. 18, 2013; revised April 2, 2014; accepted April 8, 2014.
Author contributions: A.Y. designed research; T.T., T.U., S.T., A.T., and A.Y. performed research; K.F.T., H.H., and
A.Y. contributed unpublished reagents/analytic tools; T.T., T.U., S.T., A.I., T.S.K., A.T., and A.Y. analyzed data; T.T.,
K.F.T., T.S.K., A.T., and A.Y. wrote the paper.
This work was supported by a Grant-in-Aid for Scientific Research on Innovative Areas Mesoscopic Neurocircuitry (23115103), a Grant-in-Aid for Scientific Research (B) (23300142; A.Y.), a Grant-in-Aid for Scientific Research
(C) (22500830; A.T.), and a Grant-in-Aid for Young Scientist (A) (23680042; K.T.) from the Ministry of Education,
Culture, Sports, Science and Technology of Japan; by the Prescursor Research for Embryonic Science and Technology
(PRESTO) program from Japan Science and Technology Agency (A.Y.); by Takeda Science Foundation and Kanae
Foundation by a Japan Society for Promotion of Science postdoctoral fellowship (T.T. and S.T.); and by National
Institutes of Health Grant R01 NS077408 (T.K.). We thank Dr. Y. Ootsuka and Dr. K. Matsui for assisting in data
analyses. We thank C. Saito, K. Nishimura, Y. Esaki (Nonprofit Organization Biotechnology Research and Development), and S. Nishioka (Nonprofit Organization Biotechnology Research and Development) for technical assistance.
The authors declare no competing financial interests.
This article is freely available online through the J Neurosci Author Open Choice option.
Correspondence should be addressed to Akihiro Yamanaka, PhD, Department of Neuroscience II,
Research Institute of Environmental Medicine, Nagoya University, Nagoya 464-8601, Japan. E-mail:
yamank@riem.nagoya-u.ac.jp.
DOI:10.1523/JNEUROSCI.5344-13.2014
Copyright 2014 the authors 0270-6474/14/336896-14$15.00/0

the lateral hypothalamic area (LHA) but are more numerous and
extend further rostrocaudally (Bittencourt et al., 1992; Peyron et
al., 1998; Elias et al., 2008). MCH neurons project widely
throughout the brain and densely innervate the cholinergic and
monoaminergic arousal centers (Bittencourt et al., 1992). MCH
neurons have been reported to be GABAergic (Del Cid-Pellitero
and Jones, 2012). While MCH receptor-1 (MCHR1; the only
receptor found in rodents) binding activates Gq, Gi, and Go subunits (Hawes et al., 2000), the major effect of MCHR1 binding is
to decrease cAMP levels (Chambers et al., 1999; Lembo et al.,
1999) and cellular electrophysiological studies have revealed presynaptic and postsynaptic inhibitory effects of MCH (Gao and
van den Pol, 2001; Wu et al., 2009). Thus, MCH neurons exert
neuroinhibitory effects on downstream targets.
MCH has been implicated in several functions, including
feeding, energy balance, and locomotor activity (Shimada et al.,
1998; Asakawa et al., 2002; Marsh et al., 2002; Semjonous et al.,
2009); its involvement in sleep and wakefulness has also been
extensively studied. Intracerebroventricular injections of MCH
early in the dark period increased rapid eye movement (REM)
sleep in a dose-dependent manner (Verret et al., 2003); some

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

doses also increased non-REM (NREM) sleep. A recent study


reported that optogenetic stimulation of MCH neurons increased both NREM and REM sleep (Konadhode et al., 2013).
These results are consistent with the observation that intracerebroventricular injections of MCH increase both sleep states, but
are at variance with results from functional neuroanatomical,
pharmacological (Ahnaou et al., 2008), and optogenetic (Jego et
al., 2013) studies that specifically indicate a role for MCH neurons in REM sleep regulation.
To this point, the precise role of MCH neurons in REM sleep
regulation remains unclear and the involvement of MCH neurons in NREM sleep is uncertain (Jego and Adamantidis, 2013;
Jones and Hassani, 2013; Luppi et al., 2013b; McGinty and Alam,
2013; Pelluru et al., 2013). One potential source of the discrepant
results from the optogenetic studies could be the number of
MCH neurons transfected with the light-sensitive protein and
subsequently stimulated. Therefore, we generated new transgenic
mice (Berndt et al., 2011; Han et al., 2011; Tanaka et al., 2012) to
control MCH neurons and bred this strain with three other transgenic lines to create novel bigenic strains that enabled cellspecific manipulation of MCH neurons: an E123T/T159C double
mutant of channelrhodopsin2 (ChR2) for stimulation (Berndt et
al., 2011), archaerhodopsinTP009 (ArchT) for inhibition (Han et
al., 2011), and diphtheria toxin A (DTA) for ablation. We find
that optogenetic stimulation of MCH neurons induced REM
sleep with a short latency and that continuous stimulation resulted in a net increase of REM sleep in conjunction with a decrease of NREM sleep. Although optogenetic inhibition of MCH
neurons had no effect on sleep/wakefulness, conditional ablation
of MCH neurons resulted in hyposomnia. These results demonstrate a critical role for MCH neurons in REM sleep regulation
under acute stimulation and provide unequivocal evidence for
this LHA neuronal population in sleep/wake control.

Materials and Methods


Animal usage. All experimental procedures involving animals were approved by the animal care and use committees at Nagoya University, Hokkaido University, and SRI International and were in accordance with
National Institutes of Health guidelines. All efforts were made to minimize
animal suffering or discomfort and to reduce the number of animals used.
Generation of MCH-tetracycline-controlled transactivator bacterial artificial chromosome transgenic mice. The codons of bacterial tetracycline
repressor protein and viral VP16 activator domain [tetracyclinecontrolled transactivator (tTA)] were fully mammalianized (Inamura et
al., 2012). Mouse bacterial artificial chromosome (BAC) DNA (clone
RP23-348H6) was initially modified by inserting an Rpsl-Zeo cassette (a
gift from Dr. Hisashi Mori, Toyama University) into the translation initiation site of the MCH gene followed by the replacement with a cassette
containing tTA and the SV40 polyadenylation signal. BAC DNA was
linearized by PI-SceI enzyme digestion (New England Biolabs) and injected into fertilized eggs from BDF1 mice. Two founders were obtained
and line B was used in this report.
Generation of tetracycline operator ChR2 (E123T/T159C) knock-in
mice. To achieve sufficient opsin expression (ChR2) in MCH neurons to
drive photocurrents, we used a knockin-mediated enhanced gene expression by improved tetracycline-controlled gene induction (KENGE-tet)
strategy (Tanaka et al., 2012). We used the same housekeeping gene, actb,
but the position of gene targeting was 800 bp downstream of the polyA
signal, which was different from our initial report (Tanaka et al., 2012).
We constructed a plasmid containing the tetracycline operator (TetO)
ChR2 polyA cassette with the Neo selection gene flanked on both sides by
flippase recombinase target (FRT) sites in which the following elements
were connected in tandem: TetO sequence, rabbit -globin intron,
ChR2(E123T/T159C) enhanced yellow fluorescent protein (EYFP)
cDNA, SV40 polyadenylation signal, and FRT-flanked phosphoglycerate

J. Neurosci., May 14, 2014 34(20):6896 6909 6897

kinase-EM7-Neo. To insert the above cassette downstream of the mouse


-actin gene, a 265 bp 5 homology arm consisting of the sequence from
535 to 799 bp downstream of the mouse -actin gene polyadenylation
signal (AATAAA) was connected to the 5 end of the TetO sequence, and
a 265 bp 3 homology arm consisting of the sequence from 800 to 1064 bp
downstream of the AATAAA sequence was connected to the 3 end of
Neo. To perform BAC recombination, DNA fragments with both homology arms were electroporated into bacteria carrying the BAC (clone
RP23-289L7) and the pBADTcTypeG plasmid (a gift from Dr. Manabu
Nakayama, Kazusa Institute, Japan; Nakayama and Ohara, 2005).
Kanamycin-resistant clones were selected as the modified BAC and the
TetO ChR2 polyA cassette was subsequently inserted 800 bp downstream
of mouse -actin gene AATAAA. The targeting vector was isolated from
the kanamycin-resistant, modified BAC clone using a retrieval technique
involving insertion into the pMCS-DTA plasmid (a gift from Dr. Kosuke
Yusa, Osaka University, Japan).
We subsequently obtained a targeting vector comprising a 11 kb 5
homology arm, TetO ChR2 polyA cassette with Neo, a 1.5 kb 3 homology arm, and a DTA subunit. EGR-G01 ES cell (established by F1 from
129S2 and C57BL-6-Tg(CAG/Acr-EGFP)) cells were used. We obtained
28 recombinant clones out of 96 G418-resistant clones. Recombination
was confirmed by Southern blotting with a 424 bp 3-prime outside
probe, which recognized an 8.1 kb fragment of the wild-type allele and a
6 kb fragment of the targeted allele in genomic DNA digested with XbaI.
Germline-transmitted offspring were established as TetO ChR2-Neo
knock-in mice. TetO ChR2-Neo mice were crossed with ROSA-Flp mice
(Farley et al., 2000), and FRT-flanked Neo selection markers were removed. TetO ChR2 knock-in mice were subsequently generated.
Generation of bigenic MCH mouse strains. To express ChR2, ArchT, or
DTA in MCH neurons, MCH-tTA mice were bred with TetO ChR2 mice,
TetO ArchT mice (Tsunematsu et al., 2013), or TetO DTA mice (B6.CgTg(tetO-DTA)1Gfi/J, 008468, Jackson Laboratory), respectively.
Brain slice preparation. Male and female MCH-tTA; TetO ChR2 or
MCH-tTA; TetO ArchT bigenic mice (3 6 weeks old) were used for
whole-cell recordings. The mice were deeply anesthetized with isoflurane
(Abbott Japan) and decapitated. Brains were quickly isolated in ice-cold
cutting solution consisting of (in mM) the following: 280 sucrose, 2 KCl,
10 HEPES, 0.5 CaCl2, 10 MgCl2, 10 glucose, pH 7.4 with NaOH, bubbled
with 100% O2. Brains were cut coronally into 350 m slices with a microtome (VTA-1200S, Leica). Slices containing the LHA were transferred
to an incubation chamber shielded from light and filled with a physiological
solutioncontainingthefollowing(inmM):135NaCl,5KCl,1CaCl2,1MgCl2,10
HEPES, 10 glucose, pH 7.4 with NaOH, bubbled with 100% O2. This was incubated for 1 h at room temperature (RT; 2426C).
In vitro electrophysiological recordings. At RT, the slices were transferred to a recording chamber (RC-27L, Warner Instrument) on a fluorescence microscope stage (BX51WI, Olympus). Neurons having EGFP
and EYFP fluorescence were identified as MCH neurons and were subjected to electrophysiological recordings. The fluorescence microscope
was equipped with an infrared camera (C2741-79, Hamamatsu Photonics) for infrared differential interference contrast imaging and a cooled
charge-coupled device (CCD) camera (Cascade 650, Roper Scientific) for
fluorescent imaging. Each image was displayed separately on a monitor
and saved on a computer. Recordings were performed with an Axopatch
200B amplifier (Molecular Devices) using a borosilicate pipette (GC15010, Harvard Apparatus) prepared by a micropipette puller (P-97, Sutter
Instruments), and filled with intracellular solution (4 6 M) consisting
of (in mM) the following: 138 K-gluconate, 10 HEPES, 8 NaCl, 0.2 EGTANa3, 2 MgATP, 0.5 Na2GTP, pH 7.3 with KOH. The osmolality of the
solution was checked by a vapor pressure osmometer (model 5520, Wescor). The osmolality of the internal and external solutions was 280 290
and 320 330 mOsm/l, respectively. The liquid junction potential of the
patch pipette and perfused extracellular solution was estimated to be 16
mV and was corrected in the data. Recording pipettes were under positive pressure while advancing toward individual cells in the slice, and
tight seals on the order of 1.0 1.5 G were made by negative pressure.
The membrane patch was then ruptured by suction. The series resistance
during recording was 10 25 M. The reference electrode was an AgAgCl pellet immersed in bath solution. During recordings at RT, cells

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

6898 J. Neurosci., May 14, 2014 34(20):6896 6909

were superfused with extracellular solution at a rate of 1.6 ml/min using


a peristaltic pump (Dynamax, Rainin). Blue (475 17.5 nm, 2.5 mW)
and green light (549 7.5 nm, 0.9 mW) was generated by a SPECTRA
light engine (Lumencor).
The output signal was low-pass filtered at 5 kHz and digitized at 10
kHz. Data were recorded on a computer through a Digidata 1322A
analog-to-digital converter using pClamp software version 10.2 (Molecular Devices).
Drugs. Tetrodotoxin (TTX; Wako) was dissolved in extracellular solution at a concentration of 1 M. During experiments, TTX was applied by
bath application.
Immunohistochemistry. Male and female MCH-tTA; TetO ChR2 and
MCH-tTA; TetO ArchT bigenic mice (6 weeks old) were deeply anesthetized with isoflurane and perfused sequentially with 20 ml of chilled
saline and 20 ml of chilled 10% formalin solution (Wako). The brains
were removed and immersed in the above fixative solution for 24 h at 4C
and then immersed in a 30% sucrose solution for 2 d. The brains were
quickly frozen in embedding solution (Sakura Finetechnical). For MCH
and GFP double staining, coronal sections (40 m) of MCH-tTA; TetO
ChR2 and MCH-tTA; TetO ArchT bigenic mice brains were incubated
with mouse anti-GFP antiserum (1:1000; Wako) for 24 h at 4C. These
sections were incubated with Alexa 488-labeled donkey anti-mouse IgG
(1:1000; Biotium) for 1 h at RT. The sections were then incubated with
rabbit anti-MCH antiserum (1:2000; Sigma-Aldrich) for 24 h at 4C and
incubated with Alexa 594-labeled donkey anti-rabbit IgG (1:1000; Biotium) for 1 h at RT. To confirm the specificity of antibodies, incubations
without primary antibody were conducted as a negative control in each
experiment and, in each case, no signal was observed.
To determine the number of MCH and orexin neurons, coronal sections (40 m) were stained by the avidin-biotin-peroxidase method.
Brain sections were incubated for 40 min in phosphate buffer containing
0.3% H2O2 to inactivate endogenous peroxidase. Sections were transferred into PBS containing 0.25% Triton X-100 and 1% bovine serum
albumin fraction V (PBS-BX) for 30 min and then incubated with rabbit
anti-MCH antiserum (Sigma-Aldrich) diluted 1:4000 in PBS-BX, goat
anti-orexin IgG antibody (Santa Cruz Biotechnology) diluted 1:2000 in
PBS-BX overnight at 4C. Sections were then incubated with biotinlabeled goat anti-rabbit IgG antibody (1:1000; Vector Laboratories) or
biotin-labeled horse anti-goat IgG antibody (1:1000; Vector Laboratories) for 1 h at RT followed by incubation with avidin and biotinylated
peroxidase complex solution for 1 h at RT. Bound peroxidase was visualized by DAB-buffer tablet (Merck) with 0.0015% H2O2, resulting in a
golden-brown reaction product. Approximately 10 slices from a one-infour series were selected for counting of MCH and orexin neurons.
Sections were mounted and examined with a fluorescence microscope
(BZ-9000, Keyence) or a confocal microscope (LSM710, Zeiss). Cell
counts were made bilaterally. The number of MCH neurons in the brain
is presented as the sum of the number of MCH neurons counted in every
fourth brain slice at a thickness of 40 m.
In vivo photo illumination using freely moving mice. Male MCH-tTA;
TetO ChR2 and MCH-tTA; TetO ArchT bigenic mice, 12 weeks of age or
older, were housed under controlled lighting (12 h light/dark cycle; lights
on from 8:00 to 20:00) and temperature (22 3C) conditions. Food and
water were available ad libitum. Mice were anesthetized with isoflurane
using a vaporizer for small animals (Bio Research Center) and positioned
in a stereotaxic frame (David Kopf Instruments). Plastic fiber optics (0.5
mm in diameter; Eska, Mitsubishi Rayon) were bilaterally implanted into
the hypothalamus 1 mm above the LHA (1.2 mm posterior, 1 mm
lateral from bregma, 3.5 mm depth from brain surface). Electrodes for
EEG and EMG were implanted on the skull and neck muscles, respectively. For the head-fixed mice, a U-shaped plastic plate was attached to
the skull to enable fixation of the mouses head to the stereotaxic frame
during recordings. These were attached to the skull using dental cement
(GC). The mice were then housed separately for a recovery period of 7 d.
Continuous EEG and EMG recordings were performed through a slip
ring (Air Precision, Le Pressis Robinson) designed so that the movement
of the mouse was unrestricted. EEG and EMG signals were amplified
(AB-610J, Nihon Koden), filtered (EEG, 1.530 Hz; EMG, 15300 Hz),
digitized at a sampling rate of 128 Hz, and recorded using SleepSign

software version 3 (Kissei Comtec). Blue (475 17.5 nm) and green light
(542 13.5 nm) was generated by the SPECTRA light engine and applied
through plastic optical fibers bilaterally inserted 1 mm above the LHA.
An optical swivel (COME2, Lucir) was used for unrestricted in vivo photo
illumination. The power intensities of blue and green light at the tip of
the plastic fiber optics (0.5 mm diameter) were 25.8 and 33.4 mW/mm 2,
respectively, as measured by a power meter (VEGA, Ophir Optronics).
Each animals behavior was monitored through a CCD video camera and
recorded on a computer synchronized with EEG and EMG recordings
using the SleepSign video option system (Kissei Comtec). The infrared
activity monitor was a sensor mounted on top of the cage to measure
locomotor activity. The sensors output signals (representing the magnitude of each animals movement) were digitally converted and transferred to a computer.
MCH-tTA; TetO DTA bigenic mouse surgery. MCH-tTA; TetO DTA
bigenic mice were housed under controlled lighting (12 h light/dark
cycle; lights on from 7:00 to 19:00) and temperature (22C) conditions.
Food and water were available ad libitum. Doxycycline-containing chow
(Dox chow) was made by adding 10% doxycycline (Dox) powder (Kyoritsu Seiyaku) to normal chow (Labo MR Stock, Nosan) at a final concentration of 100 mg/kg. Labo MR stock was provided during the
Dox() period (from 10 weeks of age). Mating pairs of MCH-tTA mice
and TetO DTA mice were fed with Dox-containing chow [Dox() condition] from the day of mating. During the prenatal and early postnatal
periods, Dox was supplied via maternal circulation or lactation, respectively. After weaning, MCH-tTA; TetO DTA mice were fed with Dox()
chow until the day of the experiment.
MCH-tTA; TetO DTA bigenic mice (8 weeks of age) were anesthetized
with an intraperitoneal injection of a drug mixture containing ketamine
(75 mg/kg) and medetomidine (1 mg/kg), placed in a stereotaxic device,
and chronically implanted with EEG and EMG electrodes for polysomnographic recording of sleep and wakefulness states. Two stainless-steel
screws (diameter, 1.0 mm) were implanted into the skull over the left
cerebral hemisphere (frontal: 1.5 mm lateral to midline, 1.0 mm anterior
to bregma; parietal: 1.5 mm lateral to midline, 1.0 mm anterior to
lambda) according to the atlas of Franklin and Paxinos (1997). A third
screw was placed over the frontal bone to serve as a ground electrode.
EMG activity was monitored using stainless-steel Teflon-coated wires
inserted bilaterally into the neck muscles. All electrodes were attached to
a microconnector, and the entire assembly was fixed to the skull with
dental cement. Mice were injected with indomethacin (1 mg/kg, i.p.) and
penicillin G (200 U/kg, i.m.) immediately after surgery. All procedures
were performed on a heating pad, and mice were allowed to recover for
10 14 d before being transferred to sleep recording chambers.
Vigilance state determination. Polysomnographic recordings were automatically scored offline as wakefulness, NREM sleep, or REM sleep by
SleepSign, in 4 s epochs (optogenetics experiments) or 10 s epochs (ablation experiments), according to standard criteria (Tobler et al., 1997;
Yamanaka et al., 2002). All vigilance state classifications assigned by
SleepSign were examined visually and corrected if necessary. The same
individual, blinded to genotype and experimental condition, scored all
EEG/EMG recordings. Spectral analysis of the EEG was performed by fast
Fourier transform (FFT; sampled at 128 Hz). This analysis yielded a
power spectral profile over a 0 25 Hz window with a 1 Hz resolution
divided into delta (15 Hz), theta (6 10 Hz), alpha (10 13 Hz), and beta
(1325 Hz) waves.
Statistical analysis. Data were analyzed by paired t test, unpaired t test,
one-way ANOVA, two-way ANOVA, or nonparametric KruskalWallis
test (as appropriate for the parameters examined) using KaleidaGraph
4.0 software (Hulinks). When appropriate, ANOVA tests were followed
by post hoc analysis of significance using Fishers protected least significant difference test or Bonferronis test. p values 0.05 were considered
statistically significant.

Results
Specific activation of MCH neurons using ChR2
MCH neurons in the brains of MCH-tTA; TetO ChR2 bigenic
mice specifically expressed ChR2 (Fig. 1A) as confirmed by
double-labeled immunohistochemistry. An anti-GFP antibody

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

Figure 1. Specific activation of MCH neurons. A, Schematic showing generation of bigenic MCH-tTA; TetO ChR2 mice. P, Minimal
promoter. B, C, Immunohistochemical analyses revealed that ChR2 is specifically expressed in MCH neurons in the MCH-tTA; TetO
ChR2 bigenic mouse brain. B, MCH-immunoreactive (MCH-IR) neurons (left, Alexa594, red) and ChR2::EYFP-immunoreactive
(YFP-IR) neurons (middle, Alexa488, green) located in the LHA and the zona incerta. Right, Merged image shows specific expression of ChR2 in MCH neurons. Bottom row represents higher magnifications of the regions enclosed by the squares in the top row.
Scale bars: (in top, right) top row, 200 m; (in bottom, right) bottom row, 100 m. C, Confocal microscopic image of the region
indicated by the squares in B, bottom row. Scale bar, 20 m. DH, Slice patch-clamp recordings from MCH neurons. D, Photocurrent induced by blue light (475 17.5 nm, 2.5 mW, 2 s; Vh 60 mV, TTX 1 M). E, Bar graph summarizing the data obtained
from D. F, Current-clamp mode recording, blue light pulses (2.5 mW, 10 ms, 10 Hz) generated action potentials in MCH neurons. G,
Loose cell-attached recording from MCH neurons. Blue light pulses (2.5 mW, 10 ms, 10 Hz) for 1 min every 5 min generated action
potentials. H, Summary data from G. Values represent means SEM. *p 0.05 versus EGFP-expressing MCH neurons.

J. Neurosci., May 14, 2014 34(20):6896 6909 6899

was used to detect ChR2-EYFP fusion


proteins. Merged pictures show that
ChR2-EYFP was exclusively observed in
MCH neurons in bigenic MCH-tTA; TetO
ChR2 mice (Fig. 1 B, C). Ectopic expression of ChR2 outside of MCH neurons
was not observed. Confocal microscopic
observation revealed that ChR2 was located in the plasma membrane of somata
and dendrites of MCH neurons. MCH
neurons expressing ChR2 did not show
blebbing or other features indicative of inappropriate trafficking (Fig. 1C). The
number and morphology of MCH neurons in MCH-tTA; TetO ChR2 mice were
indistinguishable from those of monogenic littermate mice (MCH-tTA or TetO
ChR2 mice, data not shown), suggesting
that ChR2 expression is not toxic to MCH
neurons. The ChR2 expression rate (YFPimmunoreactive/MCH-immunoreactive
100%) in MCH-tTA; TetO ChR2 mice
was 88.0 1.3% (n 9).
To confirm the function of ChR2 in
MCH neurons, slice patch-clamp analyses
were performed. Most MCH neurons did
not exhibit spontaneous firing (0.0 0.0
Hz, n 5). Under whole-cell voltageclamp mode, blue light (475 17.5 nm,
2.5 mW) was illuminated through the objective lens in the presence of TTX. At a
holding potential of 60 mV, blue light
illumination for 2 s induced robust inward currents (Fig. 1D). Blue light
illumination-induced transient and sustained currents were 948.2 185.5 pA
(n 9, p 0.001, ANOVA) and 335.0
64.3 pA (n 9, p 0.04, ANOVA), respectively (Fig. 1E). Under current-clamp
mode, a blue light pulse (10 ms, 10 Hz)
was applied. Blue light illumination instantaneously caused depolarization and
action potentials were generated in conjunction with light pulses (Fig. 1F ). Next,
activation of MCH neurons was confirmed using loose cell-attached recordings, which recorded firing frequencies
without affecting intracellular conditions.
Blue light illumination for 1 min (10 ms,
10 Hz) evoked repetitively induced action
potentials with high fidelity (Fig. 1G,H ).
These results strongly suggested that blue
light illumination would activate MCH
neurons in vivo.
Activation of MCH neurons increases
total time in REM sleep
To study the physiological significance of
MCH neuronal activity in the regulation
of sleep/wakefulness, MCH neurons were
activated in vivo using freely moving
MCH-tTA; TetO ChR2 mice. To determine sleep/wakefulness states, mice were

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

6900 J. Neurosci., May 14, 2014 34(20):6896 6909

Table 1. Vigilance state parameters recorded from MCH-tTA; TetO ChR2 bigenic mice, monogenic littermate mice, and MCH-tTA; TetO ArchT bigenic mice under basal
conditions
Wakefulness
REM sleep
NREM sleep

24 h
Total time (min)
Episode duration (s)
Number of episodes (counts)
Light period
Total time (min)
Episode duration (s)
Number of episodes (counts)
Dark period
Total time (min)
Episode duration (s)
Number of episodes (counts)

Control

MCH-tTA;
TetO ChR2

MCH-tTA;
TetO ArchT

Control

MCH-tTA;
TetO ChR2

MCH-tTA;
TetO ArchT

Control

MCH-tTA;
TetO ChR2

MCH-tTA;
TetO ArchT

792.5 40.7
479.1 91.7
118.0 4.6

734.5 30.3
479.3 81.9
108.1 8.1

755.9 14.8
546.7 86.9
131.4 11.4

61.5 6.8
65.8 4.8
52.6 3.4

81.2 9.3
75.1 1.6
48.7 4.6

80.8 10.6
71.7 4.7
59.4 6.6

586.0 10.6
281.6 26.2
133.3 5.1

624.2 25.1
251.6 12.1
127.0 6.8

603.2 12.3
251.2 13.2
148.2 10.5

232.4 14.7
212.3 25.9
79.4 9.0

210.2 9.2
167.3 13.7
80.9 10.4

204.9 5.9
169.8 30.7
95.2 5.7

50.5 8.1
71.5 7.6
39.5 4.3

62.5 5.0
83.7 3.7
41.3 3.8

68.1 8.1
79.8 3.1
48.2 5.1

437.1 16.5
344.8 42.8
90.4 9.8

447.3 12.2
285.7 33.9
97.7 8.9

446.9 6.2
269.0 15.4
108.6 6.1

551.0 43.9
745.8 204.8
38.6 8.2

524.4 37.6
791.3 153.4
27.3 5.9

551.0 16.8
923.6 147.3
36.2 7.2

12.7 2.9
60.2 3.4
13.1 3.0

18.7 5.7
66.4 2.9
7.4 2.0

12.7 3.4
63.6 9.7
11.2 1.7

156.4 41.3
218.3 12.6
42.9 9.4

176.9 32.2
217.4 22.4
29.3 6.5

156.3 14.9
233.4 24.6
39.6 6.4

Figure 2. In vivo optogenetic activation of MCH neurons increased time spent in REM sleep. A, F, Bar graphs illustrating the time in each vigilance state during the light period (9:00 16:00) in
(A) MCH-tTA; TetO ChR2 bigenic mice and (F ) MCH-tTA monogenic mice and TetO ChR2 monogenic mice. Blue light pulses (475 17.5 nm, 25.8 mW/mm 2, 10 ms, 10 Hz for 3 h from 10:00 to 13:00)
were applied. Blue background indicates the period of illumination. B, G, Bar graphs summarizing the data from A and F. The time spent in each vigilance state during illumination for 3 h is
summarized. Black bars, No light illumination; blue bars, blue light pulse illumination. C, H, Mean episode duration. D, I, Number of episodes. E, J, Bar graph summarizing the time in each vigilance
state for 1 h after cessation of illumination (from 13:00 to 14:00) in MCH-tTA; TetO ChR2 bigenic mice (E) and MCH-tTA monogenic mice and TetO ChR2 monogenic mice (J ). W, Wakefulness; R, REM
sleep; NR, NREM sleep. Values are represented as means SEM. *p 0.05 versus basal (no illumination).

chronically implanted with EEG and EMG electrodes and bilateral


optical fibers. The tip of the optical fiber was stereotaxically placed 1
mm above MCH neurons to illuminate ChR2-expressing MCH
neurons.
First, the vigilance state distribution of MCH-tTA; TetO ChR2
bigenic mice under basal conditions was compared with that of
littermate monogenic mice (MCH-tTA mice or TetO ChR2
mice). There were no significant differences between these strains
in the total time, mean episode duration, or the number of episodes of wakefulness, REM sleep, or NREM sleep (Table 1), indicating that ChR2 expression in MCH neurons had no effect on
sleep/wakefulness patterns.
Blue light pulses (475 17.5 nm, 25.8 mW/mm 2, 10 ms, 10
Hz) were applied during the first half of the light period (from
10:00 to 13:00) and MCH-tTA; TetO ChR2 mice without light
illumination were used as a comparison. The total time in REM
sleep was significantly increased in MCH-tTA; TetO ChR2 mice
during activation of MCH neurons (Fig. 2 A, B). During this 3 h
period, the total time spent in wakefulness, REM sleep, and

NREM sleep under basal conditions and after activation of MCH


neurons was 51.8 2.9 min (n 6) and 56.3 4.2 min [n 6,
p 0.25, not significantly different (NS) paired t test]; 15.1 1.2
min (n 6) and 46.3 2.7 min (n 6, p 0.001, paired t test);
and 113.1 3.0 min (n 6) and 77.4 3.6 min (NREM, n 6,
p 0.001, paired t test), respectively. These results indicated that
activation of MCH neuronal activity resulted in increased time
spent in REM sleep accompanied by a decrease in time spent in
NREM sleep, but without affecting the time spent in wakefulness.
The mean episode duration of NREM sleep was significantly decreased from 5.3 0.4 min (n 6) to 2.0 0.2 min (n 6, p
0.001, paired t test) by activation of MCH neurons (Fig. 2C).
Meanwhile, the number of episodes of wakefulness, REM sleep,
and NREM sleep was significantly increased (Fig. 2D). These results suggest that activation of MCH neurons increased the probability of transitions from NREM to REM sleep, resulting in
increased REM and decreased NREM sleep over the 3 h period.
Since the duration of REM episodes was unchanged, however, an
overall fragmentation of sleep/wakefulness resulted, as indicated

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

by the increased number of episodes of all states. After stimulation of MCH neurons ceased, NREM sleep was significantly increased with decreases in wakefulness and REM sleep (Fig. 2E),
suggesting a compensatory rebound in NREM sleep.
Blue light illumination (475 17.5 nm, 25.8 mW/mm 2, 10
ms, 10 Hz) had no effect on sleep/wakefulness in monogenic
MCH-tTA or TetO ChR2 mice (Fig. 2F ). Neither the total time
(Fig. 2G), nor mean episode duration (Fig. 2H ), nor the number
of episodes (Fig. 2I ) during illumination for 3 h, nor the total
time after illumination (Fig. 2J ) of each vigilance state were significantly affected by blue light pulse.
Does activation of MCH neurons induce physiological
REM sleep?
Although activation of MCH neurons increased the time spent in
REM sleep as determined from EEG and EMG recordings in
freely moving mice, it was unclear whether this state met criteria
for physiological REM sleep other than those based on EEG and
EMG recordings. To evaluate this further, eye and whisker movements during activation of MCH neurons were visually analyzed.
To monitor eye and whisker movements during MCH neuron
stimulation, mice were affixed to a stereotaxic frame via a plastic
U-shape frame fixed onto the skull. After a week adaptation and
training to these conditions, mice spontaneously cycle between
wakefulness and sleep. While a mouse was in spontaneous NREM
sleep, MCH neurons were activated and the EEG and EMG indicated signs of REM sleep shortly after initiation of stimulation
(475 17.5 nm, 25.8 mW/mm 2, 10 ms, 10 Hz). During MCH
neuron activation-induced REM sleep, the eyes and whiskers
moved intermittently, similar to that observed during spontaneously occurring REM sleep. These results suggest that MCH neuron activation-induced REM sleep is indistinguishable from
physiologically occurring REM sleep in terms of EEG, EMG, and
eye and whisker movement indicators.
Activation of MCH neurons during NREM sleep but not
during wakefulness induces REM sleep
The results presented in Figure 2 clearly demonstrate that REM
sleep can be induced by MCH neuron stimulation during NREM
sleep. To determine whether activation of MCH neurons during
wakefulness also induces a transition to REM sleep, MCH neurons were intermittently activated for 1 min (475 17.5 nm, 25.8
mW/mm 2, 10 ms, 10 Hz) every 5 min from 20:00 (the onset time
of the active dark period) to 16:46 (the second half of the inactive
light period). MCH-tTA; TetO ChR2 mice without light illumination were used as controls. Figure 3A shows a representative
hypnogram from MCH-tTA; TetO ChR2 mice under basal conditions (without light illumination). In comparison, the hypnogram of MCH-tTA; TetO ChR2 mice with blue light illumination
(Fig. 3B) shows frequent transitions to REM sleep in conjunction
with light illumination. However, transitions to REM sleep were
restricted to epochs during which mice were in NREM sleep.
Activation of MCH neurons in NREM sleep immediately decreased EEG delta power and increased theta power, suggesting a
transition to REM sleep (Fig. 3C,D). Mean REM sleep latency
calculated from the initiation of blue light illumination was
12.4 1.2 s (n 129). REM sleep was often terminated when
blue light illumination ended. In contrast, activation of MCH
neurons did not induce a transition to REM sleep when mice were
awake (Fig. 3 E, F ). EEG and EMG were indistinguishable from
before MCH neuron activation.
Figure 3GJ presents the percentage of transitions during and
after 1 min MCH neuron activation during the dark period (Fig.

J. Neurosci., May 14, 2014 34(20):6896 6909 6901

3G,H ) and in the light period (Fig. 3 I, J ). Activation of MCH


neurons was initiated either when the mice were awake (Fig.
3G,I ) or in NREM sleep (Fig. 3 H, J ). Activation of MCH neurons
while the mice were awake during either the dark or light period
did not affect vigilance states; wakefulness was maintained both
during and after activation of MCH neurons (Fig. 3G,I ). In contrast, activation of MCH neurons during NREM sleep greatly
increased the probability of transition to REM sleep (Fig. 3 H, J ).
REM sleep duration was 1 min, probably due to the duration of
light illumination. MCH neuron activation during NREM sleep
in the light period was particularly effective in facilitating the
transition from NREM to REM sleep (Fig. 3J ): the proportion of
mice transitioning from NREM sleep to REM sleep within 1 min
of MCH neuron activation was 89.7 1.2% (n 7, p 0.001,
unpaired t test). To exclude the effect of blue light illumination
alone on the transition to REM sleep, littermate monogenic
MCH-tTA or TetO ChR2 mice were subjected to blue light illumination. In these mice, the transition probability from NREM
sleep to REM sleep after 1 min blue light MCH neuron illumination was only 10.4 1.6% (n 8). The EEG power spectral
analysis by FFT showed that EEG power of 3 8 Hz was significantly increased during light illumination in the light period
(8:00 to 16:46; Fig. 3K ). The average EEG power density in the
delta (15 Hz) and theta (6 10 Hz) wave bands was significantly increased as well during light illumination (Fig. 3L),
indicating an increase in sleep intensity during activation of
MCH neurons.
Acute inhibition of MCH neurons has no effect on
sleep/wake states
Next, the effect of acute inhibition of MCH neurons on vigilance
states was determined. To strongly inhibit MCH neurons over a
long period using optogenetics, TetO ArchT mice (Tsunematsu et
al., 2013) were bred with MCH-tTA mice to generate bigenic
MCH-tTA; TetO ArchT mice (Fig. 4A).
The specific expression of ArchT in MCH neurons was confirmed by a double-label immunohistochemical study. An anti-GFP
antiserum was used to detect ArchT-EGFP fusion proteins. A merged
picture (GFP-immunoreactive and MCH-immunoreactive) revealed
that ArchT-EGFP was exclusively expressed in MCH neurons
(Fig. 4 B, C). ArchT expression was restricted to MCH neurons;
ectopic expression of ArchT outside of MCH neurons was not
observed. Confocal microscopic observation revealed that ArchT
was present in the soma and dendrites of MCH neurons. The
number and morphology of MCH-immunoreactive neurons
in MCH-tTA; TetO ArchT mice were indistinguishable from
those of monogenic littermate mice (MCH-tTA or TetO ArchT
mice; data not shown), suggesting that ArchT expression is not
toxic to MCH neurons. ArchT was expressed in 97.3 2.0%
(n 6, GFP-immunoreactive/MCH-immunoreactive 100%)
of MCH-immunoreactive neurons.
Slice patch-clamp recording from MCH neurons confirmed the
function of ArchT in MCH neurons. Under whole-cell currentclamp mode, a rectangular depolarizing current (0.2 Hz, 100 ms)
was injected through the recording electrode to generate artificial
action potentials since most MCH neurons were silent. Green light
illumination (549 7.5 nm, 0.9 mW) for 1 min completely inhibited current injection-generated action potentials (Fig. 4D). Green
light illumination completely inhibited generation of action potentials. The firing probability during green light illumination was 0.0
0.0% (n 7, p 0.001; Fig. 4E). After termination of green light
illumination, neuronal discharge rate recovered to 98.8 1.3% of
the preillumination firing rate (n 7, p 0.24, NS).

6902 J. Neurosci., May 14, 2014 34(20):6896 6909

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

Figure 3. Acute activation of MCH neurons induces transitions from NREM sleep to REM sleep. A, B, Representative hypnograms of (A) no illumination and (B) blue light illumination for 1 min
every 5 min (25.8 mW/mm 2, 10 ms, 10 Hz) from 20:00 to 16:46. The upper panel is the dark period (20:00 8:00); lower panel is the light period (8:00 20:00); blue bars indicate 1 min periods of
blue illumination. C, E, Representative traces for EEG (upper trace) and EMG (lower trace). Blue pulses were applied during NREM sleep (C) or wakefulness (E). D, F, EEG power spectra corresponding
to C and E. GJ, Graphs show percentage of animals in each vigilance state during blue pulse illumination. G, H, Dark period. I, J, Light period. Blue light illumination occurred during a waking period
in G and I and during NREM sleep in H and J. The blue lines above each graph indicate 1 min blue light illumination pulses. K, Power spectral analysis of EEG in the light period (8:00 16:46) without
illumination (black) and with illumination (blue). L, The average EEG power densities in the delta, theta, alpha, and beta wave bands in the light period (8:00 16:46). W, Wakefulness; R, REM sleep;
NR, NREM sleep; LP, light period; DP, dark period. Values are represented as means SEM. *p 0.05 versus no illumination.

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

J. Neurosci., May 14, 2014 34(20):6896 6909 6903

Figure 4. In vivo optogenetic inhibition of MCH neurons has no effect on vigilance states. A, Schematic shows generation of bigenic MCH-tTA; TetO ArchT mice. P, Minimal promoter. B, C,
Immunohistochemical analyses revealed that ArchT is specifically expressed in MCH neurons in the MCH-tTA; TetO ArchT bigenic mouse brain. B, MCH-immunoreactive (MCH-IR) neurons (left, red;
Alexa594) and ArchT::GFP-immunoreactive (GFP-IR) neurons (middle, green; Alexa488) located in the LHA and the zona incerta. Right, Merged image shows specific expression of ArchT in MCH
neurons. Bottom row are higher magnifications of the regions enclosed by the squares in the top row. Scale bars: (in top, right) top row, 200 m; (in bottom, right) bottom row, 100 m. C, Confocal
microscopic image of the region indicated by the squares in B, bottom row. Scale bar, 20 m. D, E, Slice patch-clamp recordings from MCH neurons. D, Upper trace, Current-clamp recording from
MCH neuron. Lower trace, Injected current through the recording electrode (0.2 Hz, 140 pA, 100 ms) to generate action potentials. Green light illumination (549 7.5 nm, 0.9 mW, 1 min) inhibited
firing elicited by current injection. E, Bar graph indicating the firing probability calculated from the data in D (n 7). Firing probability is normalized to the 1 min before light illumination (Pre). Green
light was applied for the duration indicated by the green bars. Values represent means SEM. *p 0.05 versus Pre. F, Green light illumination (542 13.5 nm, 33.4 mW/mm 2, 3 h from 13:00
to 16:00) in the light period. GI, Bar graphs indicating the total time in each vigilance state (G), episode duration (H ), and number of episodes (I ) during green light illumination. W, Wakefulness;
R, REM sleep; NR, NREM sleep. Values are represented as means SEM. *p 0.05 versus basal (no illumination).

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

6904 J. Neurosci., May 14, 2014 34(20):6896 6909

During the light (inactive) period, green light (542 13.5 nm,
33.4 mW/mm 2) was illuminated through fiber optics onto the
hypothalamus of MCH-tTA; TetO ArchT mice for 3 h from 13:00
to 16:00 using the same methods used for activation of MCH
neurons. MCH-tTA; TetO ArchT mice without illumination were
used as controls. Optogenetic inhibition of MCH neurons had no
significant effect on time spent in any vigilance state (Fig. 4 F, G).
In the control condition, the time spent in wakefulness, REM
sleep, and NREM sleep over the 3 h period was 47.7 3.5 min
(n 6), 18.6 2.9 min (n 6), and 113.6 2.1 min (n 6),
respectively. The time spent in wakefulness, REM sleep, and
NREM sleep during green light illumination was 46.6 5.9 min
(n 6, p 0.88, NS, paired t test), 14.5 3.3 min (n 6, p
0.054, NS, paired t test), and 118.9 5.5 min (n 6, p 0.41,
NS, paired t test), respectively. Inhibition of MCH neurons also
had no effect on episode duration (Fig. 4H ) or the number of
episodes (Fig. 4I ). As indicated in Table 1, ArchT expression in
MCH neurons had no effect on basal sleep/wakefulness patterns
(total time and episode duration of each state). These results
suggest that silencing of MCH neurons has little effect on the
initiation and maintenance of REM sleep.
Ablation of MCH neurons decreases time spent in
NREM sleep
Inhibition of MCH neurons using optogenetics had little effect
on sleep/wakefulness. However, due to the technical limitations
of optogenetic inhibition, particularly for long periods of inhibition, we wanted to induce a chronic loss of function to further
evaluate the role of MCH neurons in sleep/wakefulness regulation. Thus, we used the tet-off system mice to specifically ablate
MCH neurons. MCH-tTA mice were bred with TetO DTA mice
to generate MCH-tTA; TetO DTA bigenic mice (Fig. 5A). In these
mice, DTA expression is restricted to MCH neurons and DTA
induces cell death by inhibiting protein synthesis. The timing of
DTA expression is controlled by the presence or absence of Dox
in the chow. In the presence of DTA, tTA loses its ability to bind
the TetO sequence (Fig. 5B). However, in the absence of Dox, tTA
induces DTA expression. Here, MCH-tTA; TetO DTA mice were
fed with chow containing Dox (100 mg/kg) until 10 weeks of age.
Then, Dox() chow was replaced with Dox() chow, the timing
of which was defined as Dox() 0 week (Fig. 5C).
MCH neuron-specific and temporally controlled ablation was
confirmed by immunohistochemical studies using an anti-MCH
antibody. The number of MCH-immunoreactive neurons was
counted in C57BL/6J wild-type mice, in TetO DTA monogenic
mice at 10 weeks of age, and in MCH-tTA; TetO DTA bigenic
mice at DOX() 0, 1, 2, 3, and 4 weeks. At Dox() 0 week, the
number of MCH neurons was 1028 33 (n 3). This was
comparable to the number of MCH neurons in the C57BL/6J
wild-type mice (1076 61, n 4, p 1, KruskalWallis) and in
the TetO DTA monogenic mice (1031 65, n 3, p 1,
KruskalWallis). Meanwhile, the number of MCH neurons dramatically decreased after Dox removal (Fig. 5 D, E). The number
of MCH neurons at Dox() 1, 2, 3, and 4 weeks was 600 52
(58.3%, n 3, p 0.001, KruskalWallis), 112 18 (10.9%, n
4, p 0.001, KruskalWallis), 55 4 (5.4%, n 3, p 0.001,
KruskalWallis), and 28 9 (2.7%, n 4, p 0.001, Kruskal
Wallis), respectively. To confirm whether DTA-induced cell
death induction was restricted to MCH neurons, orexin/hypocretin neurons were stained and counted (Fig. 5 F, G). Orexin
neurons are distributed in the lateral hypothalamic area but do
not colocalize with MCH (Broberger et al., 1998; Elias et al., 1998;
Peyron et al., 1998; Bayer et al., 2002). The number of orexin

neurons was comparable to that of C57BL/6J wild-type mice and


that of TetO DTA monogenic mice (Fig. 5G). In addition, the
number of orexin neurons was unaffected by Dox removal for 4
weeks in MCH-tTA; TetO DTA mice (Fig. 5G). No signs of unhealthiness, shape, or size of orexin neurons were found even
though orexin neurons are located near MCH neurons. These
results confirm that the tet-off system functioned correctly in the
transgenic mice and that MCH neurons were specifically ablated
after removing Dox from the chow.
Next, the effect of ablation of MCH neurons on sleep/wakefulness regulation was analyzed. EEG and EMG electrodes were
implanted in MCH-tTA; TetO DTA mice at 8 weeks of age and
EEG and EMG recordings were started after 2 weeks recovery at
10 weeks of age.
MCH-tTA; TetO DTA mice, fed with Dox() chow [the
Dox() group], were used as controls. The total time in wakefulness, REM sleep, and NREM sleep was unaffected in the
Dox() group during recordings [from Dox() 0 week to
Dox() 4 weeks]. In contrast, the time spent in wakefulness was
significantly increased and the time in NREM sleep decreased in
the Dox() group (Fig. 5H ). At Dox() 4 weeks, total time in
wakefulness in the Dox() group and the Dox() group was
697.1 29.6 min (n 5) and 793.1 13.5 min (n 8, p 0.001,
unpaired t test), respectively. At Dox() 4 weeks, the total time in
NREM sleep in the Dox() group and the Dox() group was
664.2 30.5 min (n 5) and 572.7 14.8 min (n 8, p 0.01,
unpaired t test), respectively. Although increases in wakefulness
were observed during both the light and dark periods, the increase was more pronounced in the light period. Interestingly,
however, the time spent in REM sleep was unaffected in either the
light or dark period (Fig. 5He,Hf ).
The mean episode duration of NREM sleep across the 24 h
period and during the dark period significantly decreased from
Dox() 2 weeks (Fig. 5Ig,Ii). The mean episode durations of
wakefulness and REM sleep were not significantly affected (Fig.
5IaIf). These results provide evidence that MCH neurons are
involved in the regulation of NREM sleep in addition to their role
in REM sleep demonstrated above.
To confirm that the ablation of MCH neurons did not affect
normal cortical activity, cortical EEG power spectral analyses
were compared between Dox() 0 week and Dox() 4 weeks in
MCH-tTA; TetO DTA mice (Fig. 6). The power spectra in both
REM and NREM sleep of Dox() 4 weeks mice were indistinguishable from those of Dox() 0 week mice, suggesting that
ablation of MCH neurons had no effect on basal cortical activity.

Discussion
In the present study, newly developed transgenic mouse strains
enabled manipulation of the activity and fate of MCH neurons in
vivo. The results obtained support a role for MCH neurons in the
regulation of both REM and NREM sleep.
Acute and reversible manipulation of MCH neuronal activity
using optogenetics
A role for MCH neurons in sleep/wake regulation has previously
been suggested based on studies using gene knock-out mice and
pharmacological studies. Prepro-MCH gene knock-out mice exhibited longer wake bouts during the dark period but REM sleep
was unchanged (Willie et al., 2008). MCHR1 gene knock-out
mice increased wakefulness and decreased NREM sleep (Ahnaou
et al., 2011). MCHR1 antagonists decreased deep NREM and
REM sleep quantities primarily by reducing the mean episode
duration (Ahnaou et al., 2008). MCH neurons are silent during

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

J. Neurosci., May 14, 2014 34(20):6896 6909 6905

Figure 5. Ablation of MCH neurons increases wakefulness and decreases NREM sleep in MCH-tTA; TetO DTA bigenic mice. A, Schematic showing generation of bigenic MCH-tTA; TetO DTA mice. P,
Minimal promoter. B, Schematic illustration of the tetracycline-controlled gene expression system and tTA-induced DTA expression in MCH neurons. In the presence of Dox, Dox binds tTA protein and
DTA expression is repressed. C, Schematic shows the experimental protocol. In the presence of Dox, DTA expression is suppressed and MCH neurons are intact. Green bar, Chow contained Dox
[Dox()]. Open bar, Chow without Dox [Dox()]. D, Immunohistochemistry indicates MCH neuron-specific ablation in Dox(); 0, 1, 2, and 4 weeks (wk) (Figure legend continues.)

6906 J. Neurosci., May 14, 2014 34(20):6896 6909

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

waking and discharge during sleep and are


maximally active during REM sleep (Hassani et al., 2009). Despite these reports,
however, the specific role of MCH neurons in sleep/wakefulness regulation remains uncertain.
To help clarify the role of MCH neurons in sleep/wake control, we determined the effects of acute optogenetic
activation of MCH neurons. To activate
these cells, we used a ChR2 (E123T/
T159C) double mutant that is suitable for
reliable and fast optical stimulation
(Berndt et al., 2011). We found that acute
activation of MCH neurons significantly
increased the time spent in REM sleep in
association with decreased NREM sleep.
These results were in good agreement
with those of a previous study in which
ChR2 was expressed in MCH neurons using a Cre-dependent viral vector in
prepro-MCH-cre transgenic mice (Jego et
al., 2013). Interestingly, NREM rebound
was observed after stimulation of MCH
neurons. This might be caused simply by
NREM suppression during activation of
MCH neurons. Alternatively, this might
suggest that homeostatic pressure for
NREM sleep accumulates not only in
wakefulness, but also in REM sleep.
On the other hand, optogenetic activation of MCH neurons has also been reported to increase both NREM and REM
sleep (Konadhode et al., 2013). This discrepancy may be due to differential ChR2
expression among MCH neurons. The Figure 6. Ablation of MCH neurons had no effect on EEG power spectra during REM sleep and NREM sleep. AF, Power spectral
present study and that of Jego et al. (2013) analysis of EEG recorded from MCH-tTA; TetO DTA bigenic mice. Black lines indicate the presence of Dox (MCH neurons are intact);
Red lines show Dox removal for 4 weeks (MCH neurons are ablated). Inset summarizing the average EEG power density in the delta
report that 80% of MCH neurons ex- (15 Hz), theta (6 10 Hz), alpha (10 13 Hz), and beta (1325 Hz) bands during REM sleep (AC) and NREM sleep (DF). Values
pressed ChR2, whereas Konadhode et al. are represented as means SEM. *p 0.05 versus 0 week.
targeted MCH neurons that were dorsal to
the fornix (53%) with relatively little exis dependent upon interanimal replicability of each virus injecpression (20%) in the lateral and ventral hypothalamic popution. In contrast, the bigenic mouse strains used in the present
lations (Konadhode et al., 2013). Activation of a subset of MCH
study do not require viral injections and we have found interanineurons might have distinct effects on NREM sleep. Jego et al.
mal expression to be reliable and reproducible.
(2013) expressed ChR2, eNpHR3.0, or ArchT in MCH neurons
We further demonstrated that acute activation of MCH neuby viral injections into Cre transgenic mice. However, the expresrons during NREM sleep reliably triggered transitions to REM
sion rate among individual mice could differ since this technique
sleep. MCH neuron activation-induced REM sleep following
NREM sleep suggests a similarity to physiological REM sleep
control. In contrast, optogenetic silencing of MCH neurons had
4
no effect on either total REM time or the duration of REM sleep
(Figure legend continued.) indicates time after Dox(). Scale bar, 300 m. E, Bar graph
episodes. These results demonstrate that MCH neuronal activity
shows the number of MCH-immunoreactive neurons (n 3 4). F, Orexin neurons were
is sufficient to induce REM sleep, but is not necessary for REM
stained with anti-orexin antibody to produce a golden-brown color (scale bar, 300 m). G, Bar
sleep occurrence. This conclusion is consistent with the observagraph summarizing the number of orexin-immunoreactive cell bodies in the LHA. H, Line
tion that REM sleep occurs even in pontine cats in which all
graphs indicate the total time of each vigilance state. Ha, Hd, Hg, Total time in wakefulness,
REM sleep, and NREM sleep, respectively. Hb, He, Hh, Total time for 24 h during the light period
structures rostral to the brainstem have been removed (Jouvet,
in wakefulness, REM sleep, and NREM sleep. Hc, Hf, Hi, Total time during the dark period in
1962), implicating neurons in the brainstem as responsible for
wakefulness, REM sleep, and NREM sleep. I, Line graphs indicate the mean duration of each
the generation of REM sleep.
vigilance state. Ia, Id, Ig, Mean duration over 24 h in wakefulness, REM sleep, and NREM sleep,
MCH neurons may cooperatively interact with other brain
respectively. Ib, Ie, Ih, Mean duration during the light period in wakefulness, REM sleep, and
regions
to regulate REM sleep. Recent studies using cFos immuNREM sleep, respectively. Ic, If, Ii, Mean duration during the dark period in wakefulness, REM
nostaining
with retrograde tracing have revealed an interaction
sleep, and NREM sleep. Green circles, Dox() condition (MCH neurons are intact); open circles
between MCH neurons and pontine sublaterodorsal tegmental
Dox() (MCH neurons are ablated). Values are represented as means SEM. *p 0.05 versus
nucleus (SLD) neurons (Clement et al., 2012; Luppi et al., 2013a).
Dox(). LP, Light period; DP, dark period.

Tsunematsu et al. MCH Neurons and Sleep/Wake Control

Other studies indicate that REM sleep is generated by glutamatergic neurons in the SLD that are specifically active during REM
sleep (Sakai and Koyama, 1996; Boissard et al., 2002; Lu et al.,
2006; Clement et al., 2011). These studies also suggest that MCH
neurons inhibit GABAergic REM-off neurons located in the ventrolateral periaqueductal gray region during REM sleep and that
these GABAergic REM-off neurons gate the activation of the
REM-on glutamatergic neurons in the SLD. MCH neurons might
function to open this gate to initiate REM sleep during NREM
sleep. This hypothesis is supported by our results following intermittent activation of MCH neurons. Although activation of
MCH neurons during NREM sleep induced REM, activation
during wakefulness had no effect on either REM or NREM sleep.
Direct transitions from wakefulness to REM sleep do not occur under normal conditions, but it occurs in narcoleptics in
whom orexin signaling pathways are interrupted. This observation indicates that, in contrast to MCH, orexin inhibits the initiation of REM sleep. Indeed, local administration of orexin into
the locus ceruleus or the laterodorsal tegmental nucleus dramatically suppresses REM sleep without affecting NREM sleep quantity (Bourgin et al., 2000; Xi et al., 2001). Orexin neurons and
MCH neurons are coextensive in the LHA, which leads us to
imagine a functional interaction between these neurons in the
regulation of sleep/wakefulness. Our observation that MCH neuron activation during NREM sleep induces REM sleep with
higher probability in the light period compared with the dark
period might reflect this interaction. Orexin concentrations in
the CSF are higher during the dark period compared with the
light period (Yoshida et al., 2001), which might interfere with the
transition from NREM to REM sleep when MCH neurons are
activated.
Chronic and irreversible ablation of MCH neurons
To further elucidate the role of MCH neurons in sleep/wakefulness, MCH neuron-specific ablation was performed. Nearcomplete MCH neuron ablation increased the total time in
wakefulness with decreased NREM sleep. However, the total time
in REM sleep was not affected. These observations indicate that
MCH neurons also play a crucial role in the regulation of NREM
sleep. The result from this chronic disruption of MCH neurotransmission conflicted with our results demonstrating that
acute activation of MCH neurons increases total time in REM
sleep. How could such different results between chronic ablation
and acute activation of MCH neurons arise? One possible explanation is suggested by a slice patch-clamp study that revealed that
MCH exhibited an inhibitory effect on orexin neurons (Rao et al.,
2008). The orexin and MCH systems often display opposing effects on physiological functions. For example, MCH knock-out
mice were hyperactive during fasting (Willie et al., 2008), whereas
orexin neuron-ablated mice were hypoactive but did not increase
locomotor activity during fasting (Hara et al., 2001; Yamanaka et
al., 2003). Orexin neurons are thought to be active during wakefulness and silent during REM sleep (Lee et al., 2005; Mileykovskiy et al., 2005; Hassani et al., 2009), whereas MCH neurons are
active in REM sleep and inactive during wakefulness (Hassani et
al., 2009). Putative somasomatic, axosomatic, and axodendritic
contacts between orexin and MCH neurons have been observed
(Bayer et al., 2002; Guan et al., 2002). Additionally, MCHR1
knock-out mice showed an increase in glutamatergic transmission onto orexin neurons. These facts support the concept that
MCH neurons have a role in the inhibition of orexin neuronal
activity. Therefore, ablation of MCH neurons might result in a
chronic activation of orexin neurons, which might cause an in-

J. Neurosci., May 14, 2014 34(20):6896 6909 6907

crease in the total wake time and decrease in NREM sleep in MCH
neuron-ablated mice. However, acute inhibition of MCH neurons affected neither the time in wakefulness nor NREM sleep
duration. Optogenetic inhibition for 3 h might not be enough to
affect orexin neurons. Alternatively, after ablation of MCH neurons, neural network reorganization might occur, resulting in
chronic activation of orexin neurons.
MCH neurons also project to forebrain areas, such as the hippocampus and cortex (Bittencourt et al., 1992; Saito et al., 1999;
Hervieu et al., 2000). High levels of MCHR1 mRNA and MCHR1
immunoreactivity have been found in these areas (Bittencourt et
al., 1992; Saito et al., 1999; Hervieu et al., 2000). Therefore, in
addition to its role in sleep/wakefulness regulation, the release of
MCH onto cortical and hippocampal cells during REM sleep
might modulate long-term plasticity and contribute to the processing of learning and memory. Similar to the result of Konadhode et al., the increase of delta and theta power during activation
of MCH neurons might suggest the role of MCH on the cortex
(Konadhode et al., 2013). Additional studies are also required to
determine the role of MCH release during REM sleep in these
brain areas.
Together, the present results show that MCH neurons in the
hypothalamus play a significant role in the regulation of both
NREM and REM sleep. Further experiments are necessary to
identify the specific neural pathways that underlie the differential
roles that MCH neurons play in the regulation of NREM versus
REM sleep. The creation of MCH-tTA; TetO DTA mice, which
enable partial lesions of the MCH neuron population to be produced by varying the duration of Dox removal from the diet, may
be a useful tool in this regard.

Notes
Supplemental material for this article is available at https://dl.
dropboxusercontent.com/u/32320004/MCH-REM.mov. This material
has not been peer reviewed.

References
Ahnaou A, Drinkenburg WH, Bouwknecht JA, Alcazar J, Steckler T, Dautzenberg FM (2008) Blocking melanin-concentrating hormone MCH1
receptor affects rat sleep-wake architecture. Eur J Pharmacol 579:177
188. CrossRef Medline
Ahnaou A, Dautzenberg FM, Huysmans H, Steckler T, Drinkenburg WH
(2011) Contribution of melanin-concentrating hormone (MCH1) receptor to thermoregulation and sleep stabilization: evidence from MCH1
(/) mice. Behav Brain Res 218:4250. CrossRef Medline
Asakawa A, Inui A, Goto K, Yuzuriha H, Takimoto Y, Inui T, Katsuura G,
Fujino MA, Meguid MM, Kasuga M (2002) Effects of agouti-related
protein, orexin and melanin-concentrating hormone on oxygen consumption in mice. Int J Mol Med 10:523525. Medline
Bayer L, Mairet-Coello G, Risold PY, Griffond B (2002) Orexin/hypocretin
neurons: chemical phenotype and possible interactions with melaninconcentrating hormone neurons. Regul Pept 104:3339. Medline
Berndt A, Schoenenberger P, Mattis J, Tye KM, Deisseroth K, Hegemann P,
Oertner TG (2011) High-efficiency channelrhodopsins for fast neuronal stimulation at low light levels. Proc Natl Acad Sci U S A 108:7595
7600. CrossRef Medline
Bittencourt JC, Presse F, Arias C, Peto C, Vaughan J, Nahon JL, Vale W,
Sawchenko PE (1992) The melanin-concentrating hormone system of
the rat brain: an immuno- and hybridization histochemical characterization. J Comp Neurol 319:218 245. CrossRef Medline
Boissard R, Gervasoni D, Schmidt MH, Barbagli B, Fort P, Luppi PH
(2002) The rat ponto-medullary network responsible for paradoxical
sleep onset and maintenance: a combined microinjection and functional neuroanatomical study. Eur J Neurosci 16:1959 1973. CrossRef
Medline
Bourgin P, Huitron-Resendiz S, Spier AD, Fabre V, Morte B, Criado JR,
Sutcliffe JG, Henriksen SJ, de Lecea L (2000) Hypocretin-1 modulates

6908 J. Neurosci., May 14, 2014 34(20):6896 6909


rapid eye movement sleep through activation of locus coeruleus neurons.
J Neurosci 20:7760 7765. Medline
Broberger C, De Lecea L, Sutcliffe JG, Hokfelt T (1998) Hypocretin/orexinand melanin-concentrating hormone-expressing cells form distinct
populations in the rodent lateral hypothalamus: relationship to the neuropeptide Y and agouti gene-related protein systems. J Comp Neurol
402:460 474. CrossRef Medline
Chambers J, Ames RS, Bergsma D, Muir A, Fitzgerald LR, Hervieu G, Dytko
GM, Foley JJ, Martin J, Liu WS, Park J, Ellis C, Ganguly S, Konchar S,
Cluderay J, Leslie R, Wilson S, Sarau HM (1999) Melanin-concentrating
hormone is the cognate ligand for the orphan G-protein-coupled receptor
SLC-1. Nature 400:261265. CrossRef Medline
Clement O, Sapin E, Berod A, Fort P, Luppi PH (2011) Evidence that neurons of the sublaterodorsal tegmental nucleus triggering paradoxical
(REM) sleep are glutamatergic. Sleep 34:419 423. Medline
Clement O, Sapin E, Libourel PA, Arthaud S, Brischoux F, Fort P, Luppi PH
(2012) The lateral hypothalamic area controls paradoxical (REM) sleep
by means of descending projections to brainstem GABAergic neurons.
J Neurosci 32:1676316774. CrossRef Medline
Del Cid-Pellitero E, Jones BE (2012) Immunohistochemical evidence for
synaptic release of GABA from melanin-concentrating hormone containing varicosities in the locus coeruleus. Neuroscience 223:269 276.
CrossRef Medline
Elias CF, Saper CB, Maratos-Flier E, Tritos NA, Lee C, Kelly J, Tatro JB,
Hoffman GE, Ollmann MM, Barsh GS, Sakurai T, Yanagisawa M,
Elmquist JK (1998) Chemically defined projections linking the mediobasal hypothalamus and the lateral hypothalamic area. J Comp Neurol
402:442 459. CrossRef Medline
Elias CF, Sita LV, Zambon BK, Oliveira ER, Vasconcelos LA, Bittencourt JC
(2008) Melanin-concentrating hormone projections to areas involved in
somatomotor responses. J Chem Neuroanat 35:188 201. CrossRef
Medline
Farley FW, Soriano P, Steffen LS, Dymecki SM (2000) Widespread recombinase expression using FLPeR (flipper) mice. Genesis 28:106 110.
CrossRef Medline
Franklin KGB, Paxinos G (1997) The mouse brain in stereotaxic coordinates. Academic: San Diego.
Gao XB, van den Pol AN (2001) Melanin concentrating hormone depresses
synaptic activity of glutamate and GABA neurons from rat lateral hypothalamus. J Physiol 533:237252. CrossRef Medline
Guan JL, Uehara K, Lu S, Wang QP, Funahashi H, Sakurai T, Yanagizawa M,
Shioda S (2002) Reciprocal synaptic relationships between orexin- and
melanin-concentrating hormone-containing neurons in the rat lateral
hypothalamus: a novel circuit implicated in feeding regulation. Int J Obes
Relat Metab Disord 26:15231532. CrossRef Medline
Han X, Chow BY, Zhou H, Klapoetke NC, Chuong A, Rajimehr R, Yang A,
Baratta MV, Winkle J, Desimone R, Boyden ES (2011) A high-light sensitivity optical neural silencer: development and application to optogenetic control of non-human primate cortex. Front Syst Neurosci 5:18.
CrossRef Medline
Hara J, Beuckmann CT, Nambu T, Willie JT, Chemelli RM, Sinton CM,
Sugiyama F, Yagami K, Goto K, Yanagisawa M, Sakurai T (2001) Genetic ablation of orexin neurons in mice results in narcolepsy, hypophagia, and obesity. Neuron 30:345354. CrossRef Medline
Hassani OK, Lee MG, Jones BE (2009) Melanin-concentrating hormone
neurons discharge in a reciprocal manner to orexin neurons across the
sleep-wake cycle. Proc Natl Acad Sci U S A 106:2418 2422. CrossRef
Medline
Hawes BE, Kil E, Green B, ONeill K, Fried S, Graziano MP (2000) The
melanin-concentrating hormone receptor couples to multiple G proteins
to activate diverse intracellular signaling pathways. Endocrinology 141:
4524 4532. CrossRef Medline
Hervieu GJ, Cluderay JE, Harrison D, Meakin J, Maycox P, Nasir S, Leslie RA
(2000) The distribution of the mRNA and protein products of the
melanin- concentrating hormone (MCH) receptor gene, slc-1, in the central nervous system of the rat. Eur J Neurosci 12:1194 1216. CrossRef
Medline
Inamura N, Sugio S, Macklin WB, Tomita K, Tanaka KF, Ikenaka K (2012)
Gene induction in mature oligodendrocytes with a PLP-tTA mouse line.
Genesis 50:424 428. CrossRef Medline
Jego S, Adamantidis A (2013) MCH neurons: vigilant workers in the night.
Sleep 36:17831786. CrossRef Medline

Tsunematsu et al. MCH Neurons and Sleep/Wake Control


Jego S, Glasgow SD, Herrera CG, Ekstrand M, Reed SJ, Boyce R, Friedman J,
Burdakov D, Adamantidis AR (2013) Optogenetic identification of a
rapid eye movement sleep modulatory circuit in the hypothalamus. Nat
Neurosci 16:16371643. CrossRef Medline
Jones BE, Hassani OK (2013) The role of Hcrt/Orx and MCH neurons in
sleep-wake state regulation. Sleep 36:1769 1772. CrossRef Medline
Jouvet M (1962) Research on the neural structures and responsible mechanisms in different phases of physiological sleep (in French). Arch Ital Biol
100:125206. Medline
Konadhode RR, Pelluru D, Blanco-Centurion C, Zayachkivsky A, Liu M,
Uhde T, Glen WB Jr, van den Pol AN, Mulholland PJ, Shiromani PJ
(2013) Optogenetic stimulation of MCH neurons increases sleep. J Neurosci 33:1025710263. CrossRef Medline
Lee MG, Hassani OK, Jones BE (2005) Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle. J Neurosci 25:6716 6720.
CrossRef Medline
Lembo PM, Grazzini E, Cao J, Hubatsch DA, Pelletier M, Hoffert C, St-Onge
S, Pou C, Labrecque J, Groblewski T, ODonnell D, Payza K, Ahmad S,
Walker P (1999) The receptor for the orexigenic peptide melaninconcentrating hormone is a G-protein-coupled receptor. Nat Cell Biol
1:267271. CrossRef Medline
Lu J, Sherman D, Devor M, Saper CB (2006) A putative flip-flop switch for
control of REM sleep. Nature 441:589 594. CrossRef Medline
Luppi PH, Clement O, Fort P (2013a) Paradoxical (REM) sleep genesis by
the brainstem is under hypothalamic control. Curr Opin Neurobiol 23:
786 792. CrossRef Medline
Luppi PH, Peyron C, Fort P (2013b) Role of MCH neurons in paradoxical
(REM) sleep control. Sleep 36:17751776. CrossRef Medline
Marsh DJ, Weingarth DT, Novi DE, Chen HY, Trumbauer ME, Chen AS,
Guan XM, Jiang MM, Feng Y, Camacho RE, Shen Z, Frazier EG, Yu H,
Metzger JM, Kuca SJ, Shearman LP, Gopal-Truter S, MacNeil DJ, Strack
AM, MacIntyre DE et al. (2002) Melanin-concentrating hormone 1
receptor-deficient mice are lean, hyperactive, and hyperphagic and have
altered metabolism. Proc Natl Acad Sci U S A 99:3240 3245. CrossRef
Medline
McGinty D, Alam N (2013) MCH neurons: the end of the beginning. Sleep
36:17731774. CrossRef Medline
Mileykovskiy BY, Kiyashchenko LI, Siegel JM (2005) Behavioral correlates
of activity in identified hypocretin/orexin neurons. Neuron 46:787798.
CrossRef Medline
Nakayama M, Ohara O (2005) Improvement of recombination efficiency
by mutation of red proteins. Biotechniques 38:917924. CrossRef
Medline
Pelluru D, Konadhode R, Shiromani PJ (2013) MCH neurons are the
primary sleep-promoting group. Sleep 36:1779 1781. CrossRef
Medline
Peyron C, Tighe DK, van den Pol AN, de Lecea L, Heller HC, Sutcliffe JG,
Kilduff TS (1998) Neurons containing hypocretin (orexin) project to
multiple neuronal systems. J Neurosci 18:9996 10015. Medline
Rao Y, Lu M, Ge F, Marsh DJ, Qian S, Wang AH, Picciotto MR, Gao XB
(2008) Regulation of synaptic efficacy in hypocretin/orexin-containing
neurons by melanin concentrating hormone in the lateral hypothalamus.
J Neurosci 28:91019110. CrossRef Medline
Saito Y, Nothacker HP, Wang Z, Lin SH, Leslie F, Civelli O (1999) Molecular characterization of the melanin-concentrating-hormone receptor.
Nature 400:265269. CrossRef Medline
Sakai K, Koyama Y (1996) Are there cholinergic and non-cholinergic paradoxical sleep-on neurones in the pons? Neuroreport 7:2449 2453.
CrossRef Medline
Semjonous NM, Smith KL, Parkinson JR, Gunner DJ, Liu YL, Murphy KG,
Ghatei MA, Bloom SR, Small CJ (2009) Coordinated changes in energy
intake and expenditure following hypothalamic administration of neuropeptides involved in energy balance. Int J Obes (Lond) 33:775785.
CrossRef Medline
Shimada M, Tritos NA, Lowell BB, Flier JS, Maratos-Flier E (1998) Mice
lacking melanin-concentrating hormone are hypophagic and lean. Nature 396:670 674. CrossRef Medline
Tanaka KF, Matsui K, Sasaki T, Sano H, Sugio S, Fan K, Hen R, Nakai J,
Yanagawa Y, Hasuwa H, Okabe M, Deisseroth K, Ikenaka K, Yamanaka A
(2012) Expanding the repertoire of optogenetically targeted cells with
an enhanced gene expression system. Cell Rep 2:397 406. CrossRef
Medline

Tsunematsu et al. MCH Neurons and Sleep/Wake Control


Tobler I, Deboer T, Fischer M (1997) Sleep and sleep regulation in normal and prion protein-deficient mice. J Neurosci 17:1869 1879.
Medline
Tsunematsu T, Tabuchi S, Tanaka KF, Boyden ES, Tominaga M, Yamanaka A
(2013) Long-lasting silencing of orexin/hypocretin neurons using archaerhodopsin induces slow-wave sleep in mice. Behav Brain Res 255:64
74. CrossRef Medline
Verret L, Goutagny R, Fort P, Cagnon L, Salvert D, Leger L, Boissard R, Salin
P, Peyron C, Luppi PH (2003) A role of melanin-concentrating hormone producing neurons in the central regulation of paradoxical sleep.
BMC Neurosci 4:19. CrossRef Medline
Willie JT, Sinton CM, Maratos-Flier E, Yanagisawa M (2008) Abnormal
response of melanin-concentrating hormone deficient mice to fasting:
hyperactivity and rapid eye movement sleep suppression. Neuroscience
156:819 829. CrossRef Medline
Wu M, Dumalska I, Morozova E, van den Pol A, Alreja M (2009) Melaninconcentrating hormone directly inhibits GnRH neurons and blocks kiss-

J. Neurosci., May 14, 2014 34(20):6896 6909 6909


peptin activation, linking energy balance to reproduction. Proc Natl Acad
Sci U S A 106:1721717222. CrossRef Medline
Xi MC, Morales FR, Chase MH (2001) Effects on sleep and wakefulness of
the injection of hypocretin-1 (orexin-A) into the laterodorsal tegmental
nucleus of the cat. Brain Res 901:259 264. CrossRef Medline
Yamanaka A, Tsujino N, Funahashi H, Honda K, Guan JL, Wang QP, Tominaga M, Goto K, Shioda S, Sakurai T (2002) Orexins activate histaminergic neurons via the orexin 2 receptor. Biochem Biophys Res Commun
290:12371245. CrossRef Medline
Yamanaka A, Beuckmann CT, Willie JT, Hara J, Tsujino N, Mieda M, Tominaga M, Yagami K, Sugiyama F, Goto K, Yanagisawa M, Sakurai T (2003)
Hypothalamic orexin neurons regulate arousal according to energy balance in mice. Neuron 38:701713. CrossRef Medline
Yoshida Y, Fujiki N, Nakajima T, Ripley B, Matsumura H, Yoneda H, Mignot
E, Nishino S (2001) Fluctuation of extracellular hypocretin-1 (orexin A)
levels in the rat in relation to the light-dark cycle and sleepwake activities.
Eur J Neurosci 14:10751081. CrossRef Medline

Вам также может понравиться